1. Introduction
In this work, we study an initial-boundary value problem for the Navier–Stokes–Voigt (NSV) equations that model the unsteady flow of an incompressible viscoelastic fluid:
where
denotes the bounded domain of flow in
,
, with boundary
; the vector function
represents the velocity field;
p denotes the pressure;
is the viscosity coefficient;
is a length scale parameter such that
is the relaxation time of the viscoelastic fluid;
is the external forces field; and
is the initial velocity.
Note that when the NSV system becomes the incompressible Navier–Stokes equations that describe Newtonian fluid flows. If and , then we arrive at the incompressible Euler equations governing inviscid flows.
In the literature, the NSV equations are often called the Kelvin–Voigt equations or Oskolkov’s equations. The NSV model and related models of viscoelastic fluid flows have been studied extensively by different mathematicians over the past several decades starting from the pioneering papers by Oskolkov [
1,
2]. It should be mentioned at this point that Oskolkov later admitted [
3] that these works contain some errors and not all obtained results hold. In this regard, Ladyzhenskaya remarked in her note [
4] that the method of introduction of auxiliary viscosity used in [
1,
2] is incorrect under the no-slip boundary condition and explained the reasons for this. However, it is certain that the series of Oskolkov’s works played a major role in the study of the NSV equations and stimulated further research in this direction.
Let us shortly review available literature on mathematical analysis of NSV-type models. Sviridyuk [
5] established the solvability of the weakly compressible NSV equations. In [
6], the local-in-time unique solvability of problem (
1) is proved. Korpusov and Sveshnikov [
7] investigated the blowup of solutions to the NSV equations with a cubic source. Various slip problems are studied in the papers [
8,
9,
10]. Kaya and Celebi [
11] proved the existence and uniqueness of weak solutions of the so-called g-Kelvin–Voigt equations that describe viscoelastic fluid flows in thin domains. The solvability of the inhomogeneous Dirichlet problem for the equations governing a polymer fluid flow is proved in [
12]. Berselli and Spirito [
13] showed that weak solutions to the Navier–Stokes equations obtained as limits
of solutions to the NSV model are “suitable weak solutions” [
14] and satisfy the local energy inequality. Fedorov and Ivanova [
15] dealt with an inverse problem for the NSV equations. An algorithm for finding of numerical solution of an optimal control problem for the two-dimensional Kelvin–Voigt fluid flow was proposed by Plekhanova et al. [
16]. Antontsev and Khompysh [
17] established the existence and uniqueness of the global and local weak solutions to the NSV equations with p-Laplacian and a damping term. Artemov and Baranovskii [
18] proved the existence of weak solutions to the coupled system of nonlinear equations describing the heat transfer in steady-state flows of a polymeric fluid. Mohan [
19] investigated the global solvability, the asymptotic behavior, and some control problems for the NSV model with “fading memory” and “memory of length
”.
Most of the papers mentioned above deal with the study of weak (generalized) solutions to the NSV equations in the framework of the Hilbert space techniques. Therefore, it is a relevant question to prove the existence and uniqueness of strong solutions of system (
1) in a Banach space under natural conditions on the data. Another important objective is to develop convenient algorithms for finding strong solutions or their approximations. Motivated by this, in the present work, we propose the strong formulation of problem (
1) as a nonlinear evolutionary equation in suitable Banach spaces with the initial condition
. Using the Faedo–Galerkin procedure with a special basis of eigenfunctions of the Stokes operator and deriving various a priori estimates of approximate solutions in Sobolev’s spaces
and
, we construct a global-in-time strong solution of (
1), which is unique in both two-dimensional and three-dimensional domains. We also derive the energy equality that holds for strong solutions. Moreover, it is shown that, if the external forces field
is conservative, then the
-norm of the velocity field
decays exponentially as
.
2. Preliminaries
To suggest the concept of a strong solution to problem (
1), we introduce some notations, function spaces, and auxiliary results.
For vectors
and matrices
by
and
, we denote the scalar products, respectively:
Let
be a bounded domain with sufficiently smooth boundary
. By
denote the set of
functions with support contained in
. We use the standard notation for the Lebesgue spaces
,
, as well as the Sobolev spaces
,
. When it comes to classes of
-valued functions, we employ boldface letters, for instance,
It is well known that the space Sobolev is compactly embedded in .
Let us introduce the following spaces:
It is obvious that
,
, and
are Hilbert spaces with the scalar products induced by
,
, and
, respectively. However, when studying problem (
1), in the spaces
and
, it is more convenient to use the scalar products and the norms defined as follows:
Here,
is the Leray projection, which corresponds the well-known Leray (or Hodge–Helmholtz) decomposition for the vector fields in
into a divergence-free part and a gradient part (see, e.g., [
20], Chapter IV):
where the symbol ⊕ denotes the orthogonal sum and the subspace
is defined as follows
Note that the norm
is equivalent to the norm
,
.
We introduce the equivalence relation on the space by stating that if . As usual, denotes the quotient of by .
For a function
, we set
Let us define the gradient and the norm of
as follows
Using Proposition 1.2 from ([
21], Chapter I, § 1), it is easy to verify that the norm
is well defined.
The following lemmas are needed for the sequel.
Lemma 1. Suppose is a Banach space and T is a positive number. A set of the space is relatively compact if and only if:
The proof of this lemma is given in [
22].
Lemma 2. The embedding is completely continuous.
Proof. Let
be a bounded set of
. Then
with some constant
r. Clearly, this implies that the set
is bounded in
for any
.
From the Rellich–Kondrachov theorem (see, e.g., [
23], Chapter 1, Theorem 1.12.1), it follows that the space
is compactly embedded into
. Therefore, the set
is relatively compact in the space
.
By
denote the embedding operator from
into
. Taking into account inequality (
5), we get the estimate
for any function
and for any numbers
such that
where
is the operator norm of
.
Applying Lemma 1 with , we conclude that the set is relatively compact in the space . Lemma 2 is proved. □
Lemma 3. LetThen, we havefor any . Proof. The estimate (
6) is a direct consequence of the Poincaré inequality (see, e.g., [
24], Chapter II, Theorem II.5.1). □
4. Proof of Theorem 1
To prove the existence of a strong solution to problem (
1), we use the Faedo–Galerkin method with a special basis of eigenfunctions of the Stokes operator
This linear operator is invertible and
is self-adjoint and compact as a map from
into
. From the spectral theorem for self-adjoint compact operators (see, e.g., [
25], Chapter 10, Theorem 10.12), it follows that there exist sequences
and
such that
and
is an orthonormal basis of the space
.
Let
It is easily shown that
is an orthonormal basis in the space
.
Let us fix an arbitrary number
. For each fixed integer
, we would like to define the approximate solution as follows:
where
are unknown functions such that
Let us define the matrix
and the vector
by the rules:
Then, system (
12) can be rewritten in the form
where
is a known nonlinear vector function and
.
Using integration by parts, we obtain
Therefore, the matrix
is symmetric and invertible.
Applying
to the first equation of problem (
13), we obviously get
The local existence of
on an interval
is insured by the Cauchy–Peano theorem. Thus, we have a local solution
of problem (
12) on
. Below, we obtain
a priori estimates (independent of
m) for vector function
, which entail that
.
Let us assume that
satisfies system (
12). We multiply the
jth equation of (
12) by
and sum with respect to
j from 1 to
m. Since
we get
Integrating by parts the second and third terms on the left-hand side of equality (
14), we arrive at the following relation
which, in turn, gives
for any
. Further, we multiply the last equality by 2 and integrate from 0 to
t with respect to
; this yields
Taking into account (
3) and (4), we easily derive from equality (
15) that
Here and in the succeeding discussion, the symbols
,
designate positive constants that are independent of
m. Using Grönwall’s inequality, we get
Hence,
Next, by multiplying the
jth equation of (
12) with
and summing over
, we obtain
Integrating by parts the third and fourth terms on the left-hand side of the last equality, we arrive at
From here, using (
3) and Hölder’s inequality, one can obtain
whence
With the help of inequality (
16), we get
for all
. Therefore, we have
We now multiply the
jth equation of (
12) by
and sum with respect to
j from 1 to
m. Taking into account equality (
11), we get
which leads to
From this equality, with the help of Hölder’s and Young’s inequalities, we derive
Therefore, the following inequality holds
and, using estimates (
17) and (
18), we deduce that
Integrating both sides of this differential inequality with respect to
from 0 to
t, we deduce
It follows easily that
Applying Grönwall’s inequality, we obtain
This implies that
Finally, we multiply the
jth equation of (
12) by
and sum with respect to
j from 1 to
m. Bearing in mind equality (
11), we obtain
Using Hölder’s inequality, from the last equality one can derive
Clearly, this yields the estimate
Taking into account (
17)–(
19), from the last inequality, we easily obtain that
and, hence,
From estimates (
19) and (
20) and Lemma 2, it follows that there exist a subsequence
and a function
such that
converges strongly to
in the space
as
. Without loss of generality, we can assume that
On the other hand, from (
21) it follows that
Comparing the convergence results (
23) and (
24), we obtain
Integrating the
jth equation of (
12) from 0 to
s, we obtain
Integrating by parts the third and fourth terms on the left-hand side of this equality, we arrive at
Using the convergence results (
21)–(
23), we can pass to the limit
in the last equality and obtain
Applying integration by parts again, we get
Because
is a basis of
, equality (
26) remains valid if we replace
with an arbitrary vector function
from the space
, that is
From the last equality it follows that
Using the Stokes operator
, we can rewrite this equality as follows
where
is the embedding operator.
Applying the operator
to both sides of equality (
28), we get
Since
we conclude from (
29) that
Next, differentiating both sides of (
29) with respect to
s, we get
Taking into account (
30), from the last equality we deduce that
. Hence,
Next, from equality (
27) it follows that there exists an element
such that
It is readily seen that
and, consequently, we have
Letting
, from (
33) we get
Finally, differentiating both sides of (
32) with respect to
t, we arrive at
Bearing in mind (
25), (
31), (
34), and (
35), we conclude that the pair
is a strong solution to problem (
1) on the interval
. The uniqueness of a strong solution can be proved by using arguments similar to those that are presented in [
9], thus we choose to omit the details of the corresponding proof. Since
T is arbitrary, we see that
is a solution of (
1) in the sense of Definition 1.
Next, we take the
-scalar product of (
8) with the vector function
. Using integration by parts, one can easily arrive at the energy equality (
9).
The rest of the proof consists in proving inequality (
10). If there exists a function
q from the space
such that
, then we have
i. e., the total work done by external forces
is zero.
In view of (
36), the energy equality (
9) reduces to
Differentiating the last equality with respect to
t, we get
Using inequality (
6), we obtain
and, hence,
Then, by integrating (
37) with respect to
t, we derive inequality (
10). Thus, the proof of Theorem 1 is complete.